You are on page 1of 15

Article Not peer-reviewed version

Elastic Wave Propagation Control in

Porous and Finitely Deformed Locally

Resonant Nacre-Like Metamaterials

*
Umberto De Maio , Fabrizio Greco , Paolo Nevone Blasi , Andrea Pranno , Girolamo Sgambitterra

Posted Date: 20 December 2023

doi: 10.20944/preprints202312.1475.v1

Keywords: Bioinspired materials; Bloch wave analysis; Nonlinear homogenization; Band gap; Locally

resonant; Metamaterials; Instability; Finite deformations

Preprints.org is a free multidiscipline platform providing preprint service that

is dedicated to making early versions of research outputs permanently

available and citable. Preprints posted at Preprints.org appear in Web of

Science, Crossref, Google Scholar, Scilit, Europe PMC.

Copyright: This is an open access article distributed under the Creative Commons

Attribution License which permits unrestricted use, distribution, and reproduction in any

medium, provided the original work is properly cited.


Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 20 December 2023 doi:10.20944/preprints202312.1475.v1

Disclaimer/Publisher’s Note: The statements, opinions, and data contained in all publications are solely those of the individual author(s) and
contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting
from any ideas, methods, instructions, or products referred to in the content.

Article
Elastic Wave Propagation Control in Porous and
Finitely Deformed Locally Resonant Nacre-like
Metamaterials
Umberto De Maio, Fabrizio Greco *, Paolo Nevone Blasi, Andrea Pranno
and Girolamo Sgambitterra
Department of Civil Engineering, University of Calabria, 87036 (CS), Italy; umberto.demaio@unical.it;
paolo.nevoneblasi@unical.it; andrea.pranno@unical.it; Girolamo.sgambitterra@unical.it
* Correspondence: fabrizio.greco@unical.it

Abstract: Recent studies have shown that the mechanical properties of bioinspired periodic
composite materials can be strongly influenced by finite deformation effects, leading to highly
nonlinear static and dynamic behaviors at multiple length scales. For instance, in porous periodic
nacre-like microstructures, microscopic and macroscopic instabilities may occur for a given uniaxial
loading process and, as a consequence, wave attenuation properties may evolve as a function of the
microstructural evolution, giving it the designation of metamaterials. The numerical outcomes
provide new opportunities to design bioinspired soft composite metamaterials characterized by
high deformability and enhanced elastic wave attenuation capabilities given by the insertion of
voids and lead cores.

Keywords: bioinspired materials; Bloch wave analysis; nonlinear homogenization; band gap; locally
resonant; metamaterials; instability; finite deformations

1. Introduction
Over the last twenty years, as a result of the high mechanical, thermal, and electrical
performance requirements [1], advanced composite materials have been preferred over traditional
ones in many extreme engineering applications [2–4]. For example, structural members or engine
parts of space shuttles or aircraft have been designed thanks to advanced composite materials. In
addition, recently, bioinspired microstructures were used by researchers to develop innovative
materials with exceptional properties. For instance, it has been found that nacre-like materials possess
unique mechanical properties resulting from the alternating layering of soft protein and hard
aragonite platelets [5–7] and, thanks to the recent development in additive manufacturing [8–10],
many researchers started to investigate the mechanical properties of 3D printed microstructured
bioinspired materials [11–14]. Due to their complex microstructures, such bioinspired composite
materials are generally regarded as highly heterogeneous media liable to several nonlinear
phenomena such as microscopic and macroscopic instabilities caused by large deformations [15] or,
also, microscopic damage processes could occur as a result of platelets debonding from the matrix
[16]. Several studies have demonstrated that such microscopic damage phenomena are closely related
to fracture phenomena occurring at the macroscopic scale, such as delamination and crack
propagation, and that they strongly influence the dynamic response of the structure in terms of
natural frequency vibrations [17–19], thus representing the most frequent failure precursor for
advanced materials employed in extreme engineering applications. Hyperelastic constitutive laws
are commonly employed to predict with accuracy the mechanical behavior of such advanced
composite materials in a large deformation framework. To prevent numerical modeling from
becoming too expensive from the computational point of view, several advanced numerical modeling
strategies have been proposed, for example, nonlinear homogenization [20] and multiscale methods
[21]. In recent works, it has been demonstrated that bioinspired nacre-like composite materials may

© 2023 by the author(s). Distributed under a Creative Commons CC BY license.


Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 20 December 2023 doi:10.20944/preprints202312.1475.v1

be optimized to improve their mechanical performance under static loadings [22,23], and, due to their
periodic microstructures, they are also intrinsically capable of influencing elastic wave propagation.
For this reason, their vibrational response is attracting extraordinary attention [24–26] leading to the
design of new advanced bioinspired nacre-like metamaterials, which are composite materials
characterized by periodic microstructures inspired by nature. They are commonly distinguished
from traditional bioinspired materials by properties that are not found in nature [27]. While
metamaterial research has made remarkable progress over the past decade, there are still several
drawbacks. For example, frequently, their material characteristics fail to fulfill the energy dissipation
criteria essential for managing the transmission of waves and reducing noise in mechanically
challenging procedures. In recent years, the study of elastic wave propagation in nacre-like composite
metamaterials has focused on finding the optimal combination of material and geometric parameters.
It is therefore the objective of this study to improve the scientific knowledge about nacre-like
composite metamaterials, exhibiting microscopic instability under extreme loading conditions, by
proposing a lightened nacre-like composite metamaterial with hollow reinforcing platelets and lead
cores to investigate the influence of the main microscopic material and geometrical parameters
together with the addition of lead cores on the evolution of its wave attenuation properties. A brief
recap of the theoretical concepts related to the nonlinear static and dynamic response of periodic
composite materials has been reported in Section 2, while the numerical results obtained by
superimposing an elastic wave motion through the Bloch-wave technique [28–30] on a finitely
deformed configuration of the proposed lightened composite metamaterials have been reported in
Section 3. As a result of this work, we highlighted that there are great design potentials for porous
advanced bioinspired locally resonant metamaterials characterized by excellent wave absorption
properties given by the addition of lead cores and by the onset of microscopic instabilities.

2. Theoretical background
In this section, the theoretical background related to the evaluation of the homogenized
properties (Section 2.1) is reported together with the problem statement of the Bloch-wave analysis
employed to investigate the micro- and macroscopic instability conditions and the elastic wave
propagation in prestressed periodic materials (Section 2.2).

2.1. Homogenized properties in periodic media


Consider a representative volume element (RVE) of the nonlinear composite microstructure
made by an assembly of the repetitive unit cell (RUC), as reported in Figure 1, representing a
lightened nacre-like microstructure consisting of alternating stiff platelets, matrix, and voids.

Figure 1. Homogenized lightened nacre-like composite metamaterial and the corresponding


undeformed and deformed unit cell configurations.

The volume of the homogenized nacre-like composite material is denoted by V(i ) and is
enclosed by the surface ∂V(i ) on which the first Piola-Kirchhoff traction vector acts t R . All the
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 20 December 2023 doi:10.20944/preprints202312.1475.v1

quantities reported in Figure 1 with the subscript (i) are referred to the undeformed configuration.
The volume of the primitive unit cell (RUC) is identified as U ( i ) .
The undeformed and deformed RVE configurations are associated with an infinitesimal
neighborhood of a generic point X which is related to the homogenized lightened nacre-like
composite material subjected to a macroscopic gradient deformation F . The position vectors are
defined as X and x , with reference to the undeformed and the deformed RVE configuration,
respectively. The microscopic gradient deformation tensor is defined as F = ∂x / ∂X , while its
determinant J ≡ det F represents the Jacobean of the transformation giving the volume change
measure. For an easier imposition of the essential boundary condition on the RVE, the microscopic
equilibrium problem is formulated in terms of the deformation gradient tensor F and the first-Piola
Kirchhoff stress tensor TR . By assuming a sufficiently small value of the time-like parameter t
governing the monotonically increasing loading, the quantities in the incremental form can be
considered as rate quantities. Therefore, the microconstituents are characterized by the following
incrementally linear relation:

TR = C R F  , (1)

where TR , C R and F denote the rate of the first Piola-Kirchhoff stress tensor, the fourth-order
tangent moduli tensor, and the rate of the deformation gradient tensor, respectively. Here, we
consider a finitely deformed composite material composed by nearly-incompressible hyperelastic
constituents characterized by a strain energy density W (F ) . The first and the second derivatives of
W(F ) represent the stress tensor TR and the tangent moduli tensor C R , respectively:
∂W ∂ 2W
TR = , CR = . (2)
∂F ∂F 2

By considering a quasi-static loading process and the absence of volume forces, the equation of
motion in the undeformed configuration can be written in the following form:
DivTR = 0 . (3)

The averaging relationships define the macroscopic gradient deformation tensor F and the
macroscopic first Piola-Kirchhoff stress tensor TR in terms of the boundary values of the microscopic
deformation field and the nominal traction vector (in turn functions of the volume average of F
and TR ), as follows:
1 1 1
F=
V(i ) 
∂V( i )
x ⊗ n(i ) dS (i ) =
V(i )  F ( X ) dV
B( i )
(i ) −
V(i ) 
∂H ( i )
x ⊗ n(i ) dS (i ) , (4)

1 1
TR =
V(i ) 
∂V( i )
t R ⊗ X dS (i ) =
V(i ) T
B( i )
R ( X ) dV(i ) , (5)

where t R = TR n(i ) denotes the traction vector with n(i ) the outward normal at X ∈∂V(i ) and ⊗
denotes the tensor product. Moreover, B(i ) and H (i ) denote the solid and void portions of the
volume V( i ) occupied by the RVE. The microscopic deformation field x can be expressed as a sum
of two contributes, the homogeneous deformations representing the linear part FX and the
fluctuation field representing the correction part w :
x = FX + w . (6)
By substituting the expression (6) into the expression (4), we obtain that the following integral
constraint must be satisfied to make the fluctuation field w kinematically admissible:


∂V( i )
w ⊗ n(i ) dS(i ) = 0 .
(7)
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 20 December 2023 doi:10.20944/preprints202312.1475.v1

Coherently with the periodic distribution of the reinforcing platelets and void inclusions, the
above-mentioned integral constraint could be satisfied by imposing a periodic distribution of the
stress and strain fields, corresponding, from the viewpoint of the deformation field, to impose
periodic deformations on ∂V(i ) :

x+ − x− = F (X + − X − ) , (8)

while, from the viewpoint of the stress field, it corresponds to imposing antiperiodic traction on ∂V(i ) :

(TR n(i ) ) + = (−TR n(i ) ) − , (9)

where the superscripts + and − denote pairs of points located on the opposite sides of the RVE
boundaries. The variational form of the associated boundary value problem is defined in terms of the
fluctuation field w at a given macroscopic gradient deformation tensor F :

T
B( i )
R ⋅∇δ w dV(i ) = 0 ∀δ w ∈ H #1 (V(i ) ) ,
(10)

where H #1 (V(i ) ) represents the first-order Hilbert space of the vector-valued functions periodic over
V(i ) = N U (i ) , that is over all possible assembly of k N = [0, k]
N
unit cells, with N = 2 for bidimensional
problems and N = 3 for tridimensional ones, with k denoting a strictly positive integer and # denoting
the field periodicity. The variational equation (10) is consistent with a microstructural equilibrium
state characterized by antiperiodic traction on the external boundaries (9) and zero tractions on the
hole boundaries ∂H (i ) :

TR n(i ) = 0 . (11)

By imposing an incremental change in the macroscopic gradient deformation tensor F the
following incremental boundary value problem is obtained:

C [ F + ∇w ] ⋅∇δ w dV(i ) = 0 ∀δ w ∈ H #1 (V(i ) ) ,
R
(12)
B( i )

where w represents the incremental fluctuation field. By considering the following relations:
 
TR = C R [F ] , with TR = TR , (13)

the constitutive macroscopic response is then determined in terms of homogenized tangent moduli
tensor C R :
1
(F ) =  C ( X ,F ) [I + ∇w mn ] dV(i ) ,
R R hk hk
C ijhk ijmn mn (14)
V(i ) B( i )


where w hk denotes the incremental fluctuation field induced by F = I hk with I mn
hk
= δ mhδ nk .

2.2. Nonlinear static and dynamic response of periodic media


The nonlinear static and dynamic responses of a porous periodic bioinspired microstructure
subjected to uniaxial macroscopic compressive loading processes are analyzed (see also [31] for
additional information). In the former case, the microstructure static equilibrium solution path,
together with the accompanying instability analysis, is determined, while in the latter one, the
incremental wave motion problem superimposed on a finitely deformed configuration of the
microstructure equilibrium path is considered. Interrelations between the microscopic instabilities
occurring at a given uniaxial compression and wave propagation phenomena are also analyzed.
Firstly, the static response was investigated by solving the nonlinear boundary value problem
formulated on an adequately chosen RVE, adopting the homogenization theory assumptions
together with periodic boundary conditions, as reported in Figure 2. The finite deformed
configuration was obtained by imposing the following macroscopic gradient deformation tensor:
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 20 December 2023 doi:10.20944/preprints202312.1475.v1

(1− β ) 0 
F ( β ) = (1− β )e1 ⊗ e1 + (1− β )−1 e2 ⊗ e2 = 
(1− β )−1 
, (15)
 0
with e1 , e2 denoting the unit basis vectors along the direction X1 and X2, respectively and β
denoting the load parameter. By imposing that F22 = F11−1 , an acceptable approximation of the
incompressibility constraint may be established for low percentages of void volume fractions.
Subsequently, the onset of primary instabilities with short (microscopic instability) or long
(macroscopic instability) wavelength was detected by solving the following frequency domain wave
equation through the Floquet-Bloch theorem:

Div{C R [∇κ ( X )]} + ρ(i )ω 2κ ( X ) = 0 , (16)

where ρ ( i ) is the mass density in the undeformed configuration V(i ) , ω 2 are the roots of the
R
characteristic equation, C denotes the nominal tangent moduli tensor determined with reference
to a unit cell, and κ ( X ) denotes the wave function in a periodic solid based on the Floquet-Bloch
theorem which is equal to:
0
κ ( X ) = κˆ ( X )eiK ⋅X
, (17)

where κˆ( X ) denotes a wave function which is periodic on the unit cell and K 0 denotes the Bloch
wave vector defining the direction of the wave propagation and the wavelength. Primary instabilities
are detected when the lowest eigenvalue of the equation (16) first vanishes. Secondly, the dynamic
response was investigated through a Bloch-Wave analysis by evaluating the dispersion relations
associated with the proposed periodic microstructure. Specifically, the dispersion relations evolution
for increasing levels of deformation was performed by superimposing the following Bloch-Floquet
boundary conditions on the external boundaries of the unit cell on a finitely deformed configuration:

u1 |right = u1 |left ei ( K1 L + K 2 H ) u1 |upper = u1 |lower ei ( K1 L + K 2 H )


0 0 0 0

 , , (18)
i(K 0L+ K 0H ) i(K 0L+ K 0H )
u2 |right = u2 |left e 1 2 u2 |upper = u2 |lower e 1 2

where K10 and K 20 denote the components of the Bloch wave vector K 0 which are defined as a
function of a scalar parameter k ranging from 0 to 4 following the relations reported below:

 k (π L) 0 ≤ k <1 0 0 ≤ k <1
(π L) 1≤ k < 2 ( k − 1)(π H ) 1≤ k < 2
 
K10 =  , K 20 =  . (19)
(3 − k )(π L) 2≤k <3 (π H ) 2≤k <3
0 3≤ k ≤ 4 (4 − k )(π L) 3≤ k ≤ 4

Figure 2. Geometrical representation of the RVE adopted for the nonlinear static analysis together
with the imposed prescribed displacement under the assumption of periodic fluctuations.

Ranging the parameter k from 0 to 4, the Bloch-Wave vector is swapped on the external
boundaries of the first Brillouin zone allowing to determine the so-called bandgap structure of the
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 20 December 2023 doi:10.20944/preprints202312.1475.v1

periodic unit cell in terms of dispersion graph. The dispersion curves, relating the wavelength of the
propagating waves to their frequency, were analyzed to determine the evolution of the complete
bandgaps as a function of the applied strains. The numerical investigations were carried out through
a finite element discretization performed on the simulation software COMSOL Multiphysics v6.1.

3. Numerical simulations for different geometrical and material parameters


We investigated the propagation of elastic waves firstly at the undeformed configuration (Case
1) and secondly at the bifurcated state for increasing loading levels on a standard porous nacre-like
microstructure containing periodically arranged solid and hollow platelets (Case 2) and a lead-
enhanced nacre-like microstructure containing also periodically arranged lead cores (Case 3).
The geometry of the numerical model is reported in Figure 3, the unit cell length is equal to L
and its height is equal to H, in addition, two different thicknesses for the horizontal bh and vertical bv
matrix interphases were considered. The platelet's length and height are equal to Lp and Hp,
respectively. The parametric analyses were performed by varying the geometrical parameters in
terms of volume fraction vf, the hollow platelets volume fraction vf(hp), platelets aspect ratio wp= Lp /Hp,
with Lp=10µm, and interface thickness aspect ratio wb= bv / bh with bh equal to:

( −v f wb − v f w p + v 2f wb2 − 2v 2f wb w p + v 2f w 2p + 4v f wb w p ) L p
bh = . (20)
2v f wb w p

Figure 3. Investigated lightened nacre-like unit cell and associated RVE assembly with the main
geometrical parameters.

The thickness of the hollow platelets is defined by the following relation:

Hp Lp 4 H p L p v f ( hp ) + H p2 − 2 H p L p + L2p
b= + − . (21)
4 4 4
The volume fraction v f = 8 L p H p / [(2 L p + 2bv )(4 H p + 4bh )] is evaluated by considering both the
volume occupied by the platelets and voids. The hollow platelets volume fraction vf(hp) defines the
percentage of the voids with reference to the hollow platelets (vf(hp)=1 denotes full void inclusions
while vf(hp) = 0 represents full platelets), while vf(v), vf(p) and vf(l) define the volume fraction of the voids,
platelets and the lead, respectively.
In the following, the relations between the above-reported geometrical quantities are reported:

 ( L p − 2b )( H p − 2b) Lp H p
 v f ( hp ) = with b ≤ min( , )
 Lp H p 2 2
 1
 v f ( v ) = v f ( l ) = v f v hp . (22)
4
f

 1 1 
 v f ( p ) = v f +  v f  (1 − v f )
hp

 2  2 
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 20 December 2023 doi:10.20944/preprints202312.1475.v1

The mechanical behavior of the matrix, the platelets and the lead was modeled employing a neo-
Hookean hyperelastic constitutive law based on the following strain energy density function:
1 1
W = µ (tr(C ) − 3) − µ ln( J ) + λ ln( J )2 . (23)
2 2
Where µ is the initial shear modulus, C is the right elastic Cauchy-Green tensor, J is the
Jacobian while λ defines the material compressibility which is taken equal to 1000µ to model the
incompressible behavior of material phases. The initial shear modulus of the matrix phase is set equal
to µm = 0.16 MPa corresponding to Young's modulus equal to 0.5 MPa (typical of a Tango Plus 3D
printed material) and the initial shear modulus of the platelets is µm = kµm, while their material density
and Poisson's ratio are set equal to each other as ρ( p ) = ρ ( m ) = 1.145 kg / m3 and ν ( p ) = ν ( m ) = 0.49 ,
respectively. The lead cores are characterized by the following material parameters: µ (l ) = 4929MPa
, ν (l ) = 0.42 , and ρ(l ) = 11340 kg / m3 .

3.1. Case 1: Lightened nacre-like metamaterials without hollow platelets and lead cores at the undeformed
configuration
Firstly, the bandgap structure was investigated for the limit case with vf(hp) = 1 ( corresponding to
the case without hollow platelets and lead cores) at the undeformed configuration for increasing
values of volume fraction vf (i.e. from 91% to 99%) and by varying the main geometrical and material
parameters, as reported in Table 1. The highlighted areas represent the frequency ranges
corresponding to complete bandgaps for which the wave propagation is forbidden in any direction
of wave propagation. This first set of parametric investigations was conducted focusing on high
values of stiffness ratio for a frequency range equal to 0 - 150 kHz, while lower values of the stiffness
ratio will be investigated subsequently.

Table 1. Geometrical and material parameters investigated for the composite metamaterial at the
undeformed configuration.

wp 0.25 0.5 1 2 4
wb 0.25 0.5 1 2 4
k 1,000 10,000 100,000 - -

The numerical outcomes in Figure 4 show that joint aspect ratio wb = 1 (i.e. with equal thickness
between vertical and horizontal matrix joints) leads to better performance in terms of wave
propagation attenuation compared with the other values investigated.
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 20 December 2023 doi:10.20944/preprints202312.1475.v1

Figure 4. Bandgap structure at the undeformed configuration versus the volume fraction for different
values of stiffness ratio k, platelets aspect ratio wp and joints aspect ratio wb.

Generally speaking, the numerical outcomes highlighted that the complete bandgaps become
wider for increasing values of volume fraction vf from 91% to 99%, while for vf <91%, no bandgaps
were found for all the investigated parameters. In addition, we highlighted that the ranges of
frequency of the complete bandgaps become higher for increasing values of platelets aspect ratio wp
and also wider for increasing values of stiffness ratio k; in fact, the lowest investigated stiffness ratio
k=1,000 gives the worst performance in terms of wave propagation giving few and narrow bandgaps
and thus the results were not reported. Thus, the bandgap range can be tuned by varying the
geometrical parameters vf and wp considering that the wider bandgap ranges were obtained with wb
= 1, wp = 4 and the highest investigate value of k =100,000.

3.2. Case 2: Lightened nacre-like metamaterials with hollow platelets and without lead cores for increasing
levels of deformation (standard microstructure)
The previous set of preliminary investigations reported in Section 3.1 was conducted focusing
on high values of stiffness ratio, which leads to the appearance of bandgaps in a high range of
frequency (often outside the acoustic range of frequencies), making such composite metamaterials
unsuitable for acoustic applications.
In addition, in previous work by some of the authors [15], it was obtained that with increasing
values of stiffness ratio and platelets aspect ratio, the critical load factors associated with the
macroscopic instability notably decrease, leading to a higher risk of catastrophic failures due to the
onset of macroscopic instability phenomena. Subsequently, in the light of the obtained bandgap
structures reported in Figure 4, the investigation was focused on lower values of stiffness ratio k
(leading to ranges of frequency belonging to the acoustic range 20 Hz – 20 kHz), also investigating
the influence of the applied deformations together with the onset of instabilities at the microscopic
scale. A consistent set of parametric investigations was performed concerning the following
geometrical and material parameters to identify the best combination giving the widest bandgaps:
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 20 December 2023 doi:10.20944/preprints202312.1475.v1

Table 2. Geometrical and material parameters investigated for the standard metamaterial for
increasing levels of deformation.

k 1 5 10 20 30 40 50 100
wb 1 5 10 20 30 40 50 100
wp 4 5 6 7 - - - -
vf 0.5 0.55 0.6 0.65 0.7 0.75 0.8 -
vf(hp) 1 0.95 0.9 0.85 0.8 0.75 0.7 -

After some preliminary investigation, the best combination of geometrical and material
parameters, giving a wide bandgap in the undeformed configuration, was found: Lp = 10 mm, k = 20,
wb = 50, wp = 4, vf = 0.8, vf(hp) = 0.8. First of all, the microscopic and the macroscopic instability analyses
were performed, and we found that the microscopic instability occurs before the macroscopic one at
a load parametric value equal to 0.09789, representing a percentage of deformation along the
compression direction (X1 direction with reference Figure 3) equal to about 10%. Subsequently, as can
be seen in Figure 5a, by adding a geometrical imperfection in the form of the identified critical mode
shape, the bifurcated solution was determined, and the bandgap analysis was performed at each
loading step to determine the evolution of the dispersion graphs. The proof that the primary
instability is characterized by a local instability (wavelength of the critical mode shape comparable
with the unit cell size) was given by the local instability check reported in Figure 5b.

Figure 5. Deformed configuration of the metamaterial RVE for the principal solution path step
together with the bifurcated one a) and the results of the local instability check b).

In Figure 6, the evolution of the bandgap structure at the undeformed configuration and for
increasing levels of deformation after the onset of the microscopic instability is reported. We observed
that at the undeformed configuration, a complete wide bandgap is found in the range from 2.6 kHz
to 3.2 kHz, while, as the level of deformation increases, the dispersion graph evolves considerably,
showing a thinning of the bandgaps up to the complete extinction of the one identified at the
undeformed configuration. However, new and wider complete bandgaps also appear in lower ranges
of frequencies, and numerous partial bandgaps along the X and Y directions of wave propagation
appear due to the prestress state inducing the onset of microscopic instability.
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 20 December 2023 doi:10.20944/preprints202312.1475.v1

10

Figure 6. Dispersion relations for the standard microstructure at the undeformed configuration a) and
at the bifurcated one b) with a uniaxial stretch ratio along the X direction equal to 0.3.

The results obtained through the Bloch-wave analyses were also validated by the attenuation test and
the transient analyses reported in Figure 7. In this figure, it is highlighted that the elastic pressure
waves propagating with a frequency equal to 2 kHz (outside the bandgap range) can propagate
through the proposed microstructured metamaterials while with a frequency equal to 3 kHz (inside
the complete bandgap range) the elastic wave propagation is completely forbidden.

Figure 7. Transmittance spectra at the undeformed configuration and transient simulation results for
propagating waves inside and outside the bandgap range.

Further numerical validations by the transmittance spectra determination were performed on


every investigated microstructure reported, both at the deformed and undeformed configuration, but
for the sake of brevity, only the first validation was reported in the text.
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 20 December 2023 doi:10.20944/preprints202312.1475.v1

11

3.3 Case 3: Lightened nacre-like metamaterials with hollow platelets and lead cores for increasing levels of
deformation (lead-enhanced microstructure)
In this section, the numerical results related to the evolution of the bandgap structure in a locally
resonant metamaterial were reported. The previously investigated nacre-like microstructure with
hollow platelets and voids was modified with the addition of lead cores. In Figure 8, the dispersion
relations at the undeformed and deformed configurations were reported to investigate the intricate
interplay between structural modifications and the resulting acoustic properties.

Figure 8. Dispersion relations for the lead-enhanced microstructure at the undeformed configuration
a) and at the bifurcated one b) with a uniaxial stretch ratio along the X direction equal to 0.3.

At the undeformed configuration, Figure 8a illustrates the presence of wide and complete
bandgaps spanning the frequency ranges from 2 to 3 kHz and from 3.4 kHz to 4.2 kHz, together with
two thinner bandgaps along the X2 direction of elastic wave propagation. The bandgaps found results
to be incredibly wider compared with those obtained for a microstructure without the insertion of
lead cores (Figure 7a).
The introduction of lead cores surrounded by a soft matrix plays a fundamental role in shaping
these acoustic properties. The lead cores, acting as local resonators, are influential in enhancing the
material's ability to absorb energy. This is especially noteworthy at lower vibration frequencies,
where the lead-modified microstructure demonstrates superior performance compared to its non-
modified microstructure. This highlights the potential of the locally resonant metamaterial to
attenuate vibrations within this specific frequency band effectively.
However, as deformation levels intensify, the dispersion graph undergoes significant
transformations. Notably, in Figure 7b, there is a discernible widening of the first complete bandgap,
with a complete extinction of the second complete bandgap, which is followed by the appearance of
a thinner complete bandgap in low and high-frequency ranges. In addition, the results highlight the
emergence of a new wider bandgap along the X1 direction in an incredibly low-frequency range (0.5
kHz to 1 kHz). These phenomena are directly attributed to the prestress state induced by the
structural modifications, leading to the onset of microscopic instability. With its lead-enhanced
configuration, the proposed locally resonant metamaterial offers a valuable design alternative to
control and tune the elastic wave propagation in bioinspired nacre-like metamaterials.

4. Conclusions
In this work, the main goal was to design soft metamaterials inspired by nacre, capable of
bearing significant deformations before the onset of microscopic instabilities, together with the
capability to attenuate the elastic wave propagation in specified frequency ranges beneficial for
advanced engineering applications. To achieve this goal, a comprehensive set of parametric analyses
was conducted, systematically varying the main geometrical and material parameters of a novel
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 20 December 2023 doi:10.20944/preprints202312.1475.v1

12

microstructure inspired by nacre-like and modified by the addition of reinforcing hollow platelets
and lead cores. The aim was to identify the optimal combination that would yield superior wave
attenuation capabilities.
The numerical findings highlighted that an elevated shear stiffness ratio between the reinforcing
platelets and the soft matrix contributes to the appearance of wide and complete bandgaps.
Interestingly, we found that the promising attenuation properties are not exclusive to scenarios with
high contrast between the platelets and the matrix. Instead, they can also be achieved at lower
contrast levels by tuning the percentage values of platelet, lead cores, and void inclusions volume
fraction.
Additionally, it has been observed that the initiation of microscopic instability results in a
transformation of the microstructural pattern. This transformation, in conjunction with applied
deformations, significantly influences the investigated microstructure's wave propagation
properties, giving rise to the formation of both new complete and partial bandgaps.
Definitively, this study presents novel opportunities for the design of bioinspired soft composite
metamaterials. Incorporating lead cores, acting as local resonators, we found a strong increase in the
wave attenuation properties, contributing to the overall versatility and performance of the
engineered locally resonant metamaterials. Consequently, this research provides valuable insights
for developing innovative, bioinspired metamaterials tailored for applications requiring high
deformability and effective wave attenuation capabilities.

Author Contributions: Conceptualization: A.P.; methodology: F.G. and A.P.; software: U.D.M., G.S. and A.P.;
validation: U.D.M., G.S. and A.P.; formal analysis: U.D.M., G.S. and A.P.; investigation: U. D. M. and A.P.;
resources: U.D.M. and A.P.; data curation: U.D.M. and A.P.; writing—original draft preparation: U.D.M., G.S.
and A.P.; writing—review and editing: F.G. and P.N.B.; visualization, U.D.M. and A.P.; supervision: F.G. and
P.N.B.; project administration: F.G. and P.N.B.; funding acquisition: F.G. and P.N.B. All authors have read and
agreed to the published version of the manuscript.
Funding: Fabrizio Greco, Umberto De Maio, Paolo Nevone Blasi, Girolamo Sgambitterra gratefully
acknowledges financial support by the Next Generation EU - Italian NRRP, Mission 4, Component 2, Investment
1.5, call for the creation and strengthening of 'Innovation Ecosystems', building 'Territorial R&D Leaders'
(Directorial Decree n. 2021/3277) - project Tech4You - Technologies for climate change adaptation and quality of
life improvement, n. ECS0000009; Andrea Pranno gratefully acknowledges financial support from the POR
Calabria FESR-FSE 2014-2020, Rep. N. 1006 of 30/03/2018, Line B, Action 10.5.12. This work reflects only the
authors’ views and opinions, neither the Ministry for University and Research nor the European Commission
can be considered responsible for them.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Zhang, Q.; Cherkasov, A. V.; Xie, C.; Arora, N.; Rudykh, S. Nonlinear Elastic Vector Solitons in Hard-
Magnetic Soft Mechanical Metamaterials. International Journal of Solids and Structures 2023, 280, 112396.
https://doi.org/10.1016/j.ijsolstr.2023.112396.
2. Amarante dos Santos, F.; Fraternali, F. Novel Magnetic Levitation Systems for the Vibration Control of
Lightweight Structures and Artworks. Structural Contr & Hlth 2022, 29 (8). https://doi.org/10.1002/stc.2973.
3. Santos, F. A.; Caroço, C.; Amendola, A.; Miniaci, M.; Fraternali, F. 3D TENSEGRITY BRACES WITH
SUPERELASTIC RESPONSE FOR SEISMIC CONTROL. Int J Mult Comp Eng 2022, 20 (5), 53–64.
https://doi.org/10.1615/IntJMultCompEng.2022041968.
4. Ammendolea, D.; Greco, F.; Leonetti, L.; Lonetti, P.; Pascuzzo, A. A Numerical Failure Analysis of Nano-
Filled Ultra-High-Performance Fiber-Reinforced Concrete Structures via a Moving Mesh Approach.
Theoretical and Applied Fracture Mechanics 2023, 103877. https://doi.org/10.1016/j.tafmec.2023.103877.
5. Zhao, N.; Wang, Z.; Cai, C.; Shen, H.; Liang, F.; Wang, D.; Wang, C.; Zhu, T.; Guo, J.; Wang, Y.; Liu, X.;
Duan, C.; Wang, H.; Mao, Y.; Jia, X.; Dong, H.; Zhang, X.; Xu, J. Bioinspired Materials: From Low to High
Dimensional Structure. Adv. Mater. 2014, 26 (41), 6994–7017. https://doi.org/10.1002/adma.201401718.
6. Zhang, C.; Mcadams, D. A.; Grunlan, J. C. Nano/Micro-Manufacturing of Bioinspired Materials: A Review
of Methods to Mimic Natural Structures. Adv. Mater. 2016, 28 (30), 6292–6321.
https://doi.org/10.1002/adma.201505555.
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 20 December 2023 doi:10.20944/preprints202312.1475.v1

13

7. Bosia, F.; Dal Poggetto, V. F.; Gliozzi, A. S.; Greco, G.; Lott, M.; Miniaci, M.; Ongaro, F.; Onorato, M.;
Seyyedizadeh, S. F.; Tortello, M.; Pugno, N. M. Optimized Structures for Vibration Attenuation and Sound
Control in Nature: A Review. Matter 2022, 5 (10), 3311–3340. https://doi.org/10.1016/j.matt.2022.07.023.
8. Huang, Y.; Li, J.; Chen, W.; Bao, R. Tunable Bandgaps in Soft Phononic Plates with Spring-Mass-like
Resonators. International Journal of Mechanical Sciences 2019, 151, 300–313.
https://doi.org/10.1016/j.ijmecsci.2018.11.029.
9. Liu, F.; Li, T.; Jia, Z.; Wang, L. Combination of Stiffness, Strength, and Toughness in 3D Printed Interlocking
Nacre-like Composites. Extreme Mechanics Letters 2020, 35, 100621.
https://doi.org/10.1016/j.eml.2019.100621.
10. Slesarenko, V.; Kazarinov, N.; Rudykh, S. Distinct Failure Modes in Bio-Inspired 3D-Printed Staggered
Composites under Non-Aligned Loadings. Smart Mater. Struct. 2017, 26 (3), 035053.
https://doi.org/10.1088/1361-665X/aa59eb.
11. Wang, B.; Hu, X.; Lu, P. Modelling and Testing of Large-Scale Masonry Elements under Three-Point
Bending – Tough and Strong Nacre-like Structure Enlarged by a Factor of 20,000. Engineering Fracture
Mechanics 2020, 229, 106961. https://doi.org/10.1016/j.engfracmech.2020.106961.
12. Wei, Z.; Xu, X. Gradient Design of Bio-Inspired Nacre-like Composites for Improved Impact Resistance.
Composites Part B: Engineering 2021, 215, 108830. https://doi.org/10.1016/j.compositesb.2021.108830.
13. Wan, H.; Leung, N.; Algharaibeh, S.; Sui, T.; Liu, Q.; Peng, H.-X.; Su, B. Cost-Effective Fabrication of Bio-
Inspired Nacre-like Composite Materials with High Strength and Toughness. Composites Part B: Engineering
2020, 202, 108414. https://doi.org/10.1016/j.compositesb.2020.108414.
14. Bouville, F. Strong and Tough Nacre-like Aluminas: Process–Structure–Performance Relationships and
Position within the Nacre-Inspired Composite Landscape. J. Mater. Res. 2020, 35 (8), 1076–1094.
https://doi.org/10.1557/jmr.2019.418.
15. Greco, F.; Leonetti, L.; De Maio, U.; Rudykh, S.; Pranno, A. Macro- and Micro-Instabilities in Incompressible
Bioinspired Composite Materials with Nacre-like Microstructure. Composite Structures 2021, 269, 114004.
https://doi.org/10.1016/j.compstruct.2021.114004.
16. Greco, F.; Leonetti, L.; Lonetti, P. A Novel Approach Based on ALE and Delamination Fracture Mechanics
for Multilayered Composite Beams. Composites Part B: Engineering 2015, 78, 447–458.
https://doi.org/10.1016/j.compositesb.2015.04.004.
17. Pranno, A.; Greco, F.; Lonetti, P.; Luciano, R.; De Maio, U. An Improved Fracture Approach to Investigate
the Degradation of Vibration Characteristics for Reinforced Concrete Beams under Progressive Damage.
International Journal of Fatigue 2022, 163, 107032. https://doi.org/10.1016/j.ijfatigue.2022.107032.
18. De Maio, U.; Gaetano, D.; Greco, F.; Lonetti, P.; Nevone Blasi, P.; Pranno, A. The Reinforcing Effect of Nano-
Modified Epoxy Resin on the Failure Behavior of FRP-Plated RC Structures. Buildings 2023, 13 (5), 1139.
https://doi.org/10.3390/buildings13051139.
19. De Maio, U.; Gaetano, D.; Greco, F.; Lonetti, P.; Pranno, A. The Damage Effect on the Dynamic
Characteristics of FRP-Strengthened Reinforced Concrete Structures. Composite Structures 2023, 309, 116731.
https://doi.org/10.1016/j.compstruct.2023.116731.
20. Bertoldi, K.; Bigoni, D.; Drugan, W. J. Nacre: An Orthotropic and Bimodular Elastic Material. Composites
Science and Technology 2008, 68 (6), 1363–1375. https://doi.org/10.1016/j.compscitech.2007.11.016.
21. Greco, F.; Leonetti, L.; Lonetti, P.; Luciano, R.; Pranno, A. A Multiscale Analysis of Instability-Induced
Failure Mechanisms in Fiber-Reinforced Composite Structures via Alternative Modeling Approaches.
Composite Structures 2020, 251, 112529. https://doi.org/10.1016/j.compstruct.2020.112529.
22. Grossman, M.; Pivovarov, D.; Bouville, F.; Dransfeld, C.; Masania, K.; Studart, A. R. Hierarchical
Toughening of Nacre-Like Composites. Adv Funct Materials 2019, 29 (9), 1806800.
https://doi.org/10.1002/adfm.201806800.
23. Flores-Johnson, E. A.; Shen, L.; Guiamatsia, I.; Nguyen, G. D. Numerical Investigation of the Impact
Behaviour of Bioinspired Nacre-like Aluminium Composite Plates. Composites Science and Technology 2014,
96, 13–22. https://doi.org/10.1016/j.compscitech.2014.03.001.
24. Chen, Y.; Wang, L. Multiband Wave Filtering and Waveguiding in Bio-Inspired Hierarchical Composites.
Extreme Mechanics Letters 2015, 5, 18–24. https://doi.org/10.1016/j.eml.2015.09.002.
25. Lu, Y.; Huang, G.-Y.; Wang, Y.-F.; Wang, Y.-S. A Mechanical Model for Elastic Wave Propagation in Nacre-
Like Materials With Brick-and-Mortar Microstructures. Journal of Applied Mechanics 2022, 89 (9), 091002.
https://doi.org/10.1115/1.4054897.
26. Pranno, A.; Greco, F.; Leonetti, L.; Lonetti, P.; Luciano, R.; De Maio, U. Band Gap Tuning through
Microscopic Instabilities of Compressively Loaded Lightened Nacre-like Composite Metamaterials.
Composite Structures 2022, 282, 115032. https://doi.org/10.1016/j.compstruct.2021.115032.
27. Mazzotti, M.; Foehr, A.; Bilal, O. R.; Bergamini, A.; Bosia, F.; Daraio, C.; Pugno, N. M.; Miniaci, M. Bio-
Inspired Non Self-Similar Hierarchical Elastic Metamaterials. International Journal of Mechanical Sciences
2022, 107915. https://doi.org/10.1016/j.ijmecsci.2022.107915.
Preprints (www.preprints.org) | NOT PEER-REVIEWED | Posted: 20 December 2023 doi:10.20944/preprints202312.1475.v1

14

28. Li, J.; Slesarenko, V.; Rudykh, S. Auxetic Multiphase Soft Composite Material Design through Instabilities
with Application for Acoustic Metamaterials. Soft Matter 2018, 14 (30), 6171–6180.
https://doi.org/10.1039/C8SM00874D.
29. Shim, J.; Wang, P.; Bertoldi, K. Harnessing Instability-Induced Pattern Transformation to Design Tunable
Phononic Crystals. International Journal of Solids and Structures 2015, 58, 52–61.
https://doi.org/10.1016/j.ijsolstr.2014.12.018.
30. Dalklint, A.; Wallin, M.; Bertoldi, K.; Tortorelli, D. Tunable Phononic Bandgap Materials Designed via
Topology Optimization. Journal of the Mechanics and Physics of Solids 2022, 163, 104849.
https://doi.org/10.1016/j.jmps.2022.104849.
31. De Maio, U.; Greco, F.; Luciano, R.; Sgambitterra, G.; Pranno, A. Microstructural Design for Elastic Wave
Attenuation in 3D Printed Nacre-like Bioinspired Metamaterials Lightened with Hollow Platelets.
Mechanics Research Communications 2023, 128, 104045. https://doi.org/10.1016/j.mechrescom.2023.104045.

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those
of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s)
disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or
products referred to in the content.

You might also like